0% found this document useful (0 votes)
41 views66 pages

Sara Abid

Uploaded by

Rabnawaz
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
41 views66 pages

Sara Abid

Uploaded by

Rabnawaz
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 66

i

Facile Synthesis and Electrochemical Study of TiO2/MoS2/SnS


as a High-Performance Supercapacitor Electrode Material

By

Sara Abid

(Registration No: 00000365350)

A thesis submitted to the National University of Sciences and Technology, Islamabad,

in partial fulfillment of the requirements for the degree of

Master of Science in
Chemistry

Supervisor: Dr. Adil Mansoor

Co-supervisor: Dr. Ghulam Ali

School of Natural Sciences

National University of Sciences & Technology (NUST)

Islamabad, Pakistan
(2

i
i
DEDICATION

This thesis is dedicated to my parents and siblings

For their endless love, support and encouragement

i
ACKNOWLEDGEMENTS

I’m deeply obliged to Allah ALMIGHTY for HIS countless blessings and guidance in all

stages of my life.

I would like to express my gratitude to my supervisor, Dr. Adil Mansoor, for his

continuous support and encouragement throughout my research. I’m profoundly grateful

to my co-supervisor, Dr. Ghulam Ali, for his invaluable guidance, dedication, insightful

suggestions, constructive feedback, and I’m glad I got the opportunity to work under his

supervision in USPCASE-E which led to the successful completion of my research work.

I’m thankful for my GEC members, Dr. Mudassir Iqbal and Dr. Azhar Mehmood, for their

guidance and assistance.

I also want to acknowledge the lab engineers, seniors, and lab fellows of both AEMS lab

and AIMS lab in USPCAS-E and SNS.

I’m thankful to my friends who have been there in tough times and motivated me

throughout my research.

ii
TABLE OF CONTENTS

ACKNOWLEDGEMENTS II

TABLE OF CONTENTS III

LIST OF TABLES V

LIST OF FIGURES VI

LIST OF SYMBOLS, ABBREVIATIONS AND ACRONYMS VIII

ABSTRACT IX

CHAPTER 1: INTRODUCTION 1
1.1 Supercapacitors 1
1.1.1 Electrical double layer capacitors (EDLCs): 2
1.1.2 Pseudocapacitors 3
1.1.3 Hybrid Capacitors 4
1.2 Supercapacitors based on Transition Metal Chalcogenides 5
1.2.1 Low conductivity 5
1.2.2 Poor cycle stability 6
1.2.3 Limited surface area 6
1.2.4 Aggregation and restacking issues 6
1.3 Modified transition metal chalcogenides 7
1.3.1 Carbon additives 7
1.3.2 Conductive polymers 7
1.3.3 Heterostructure formation 7
1.4 Objectives 9

CHAPTER 2: LITERATURE REVIEW 10

CHAPTER 3: EXPERIMENTAL SECTION 16


3.1 Chemicals 16

iii
3.2 Experimental 16
3.2.1 Synthesis of TiO2 16
3.2.2 TiO2/MoS2 binary composite (TM) 17
3.2.3 TiO2/SnS binary composite (TS) 18
3.2.4 TiO2/MoS2/SnS (TMS) 19
3.3 Material characterization 20
3.4 Preparation of the electrode material 21

CHAPTER 4: RESULTS AND DISCUSSIONS 23


4.1 FTIR analysis 23
4.2 X-ray diffraction (XRD) analysis 24
4.3 X-ray photoelectron spectroscopy (XPS) analysis 26
4.4 SEM and EDS analysis 29
4.5 Thermogravimetric analysis (TGA) 30
4.6 BET analysis 32
4.7 Cyclic Voltammetry (CV) analysis 33
4.8 Galvanostatic charge-discharge (GCD) analysis 37
4.9 EIS analysis 40
4.10 Ex-situ XRD and SEM Analysis 43
4.11 Full Cell Testing 44

CHAPTER 5: CONCLUSION 47

REFERENCES

iv
LIST OF TABLES

Page No.

Table 4.1: Analysis of surface area and BJH pore data of TM, TS, and TMS
composites. ................................................................................................................ 32
Table 4.2: Specific capacitances of TM, TS, and TMS electrodes calculated from CV.
................................................................................................................................... 34
Table 4.3: Specific capacitances of TM, TS, and TMS at varying current densities.
................................................................................................................................... 38
Table 4.4: Rs and Rct values of TM, TS and TMS electrodes. .................................. 41
Table 4.5: Specific capacitances of similar electrodes as reported in literature. ...... 42
Table 4.6: Specific Capacitance of full cell 0.5-10 A/g. ........................................... 45

v
LIST OF FIGURES

Page No.

Figure 1.1: Energy capability and power output [3]. .......................................................... 1


Figure 1.2: The charge storage mechanism of EDLC [7]. .................................................. 3
Figure 1.3: Pseudocapacitors exhibiting adsorption and faradaic reactions [7]. ................ 4
Figure 1.4: Types of supercapacitors. ................................................................................. 5
Figure 1.5: An illustration of the modification strategies of TMCs for superior
electrochemical performance [21]. ..................................................................................... 9
Figure 3.1: Synthesis Scheme of TiO2 .............................................................................. 17
Figure 3.2: Synthesis Scheme of TiO2/MoS2.................................................................... 18
Figure 3.3: Synthesis Scheme of TiO2/SnS ...................................................................... 19
Figure 3.4: Synthesis Scheme of TiO2/MoS2/SnS ............................................................ 20
Figure 4.1: FTIR of TM, TS and TMS composites. ......................................................... 24
Figure 4.2: XRD pattern of TM, TS and TMS composites. ............................................. 26
Figure 4.3: Survey Spectra of TM, TS and TMS composites. .......................................... 27
Figure 4.4: (a) S, (b) O, (c) Ti, (d) Mo, and (e) Sn, XPS spectra respectively. ................ 29
Figure 4.5: (a, d) TM, (b, e) TS and (c, f) TMS, SEM images. ........................................ 30
Figure 4.6: EDS Spectra and elemental of TMS composite. ............................................ 30
Figure 4.7: TGA of TM, TS and TMS composites. .......................................................... 31
Figure 4.8: (a) BET sorption isotherms, (b) BJH pore data of TM, TS and TMS composites.
........................................................................................................................................... 33
Figure 4.9: CV profiles between 5–100 mV/s of (a) TM, (b) TS, and (c) TMS electrodes.
(d) CV curves compared at 5 mV/s of all electrodes. (e) The capacitance values of
electrodes at all scan rates. (f) Electrochemical active surface area (ESCA) of all three
electrodes. ......................................................................................................................... 35

vi
Figure 4.10: Capacitive and diffusion contribution ratios between 5-100 mV/s for (a) TM,
(b) TS, and (c) TMS electrodes. A comparison of capacitive and total current at 40 mV/s
for (d) TM, (e) TS and (f) TMS electrodes. ...................................................................... 37
Figure 4.11: GCD curves between 1–10 A/g of (a) TM, (b) TS, and (c) TMS electrodes.
(d) Comparative GCD curves at 1 A/g. (e) Specific capacitances derived from GCD curves
at all current densities. (f) Ragone plot for the three electrodes. ...................................... 40
Figure 4.12: (a) Stability test of TMS for 10,000 cycles at 10 A/g, and (b) EIS of TM, TS,
and TMS electrodes. ......................................................................................................... 42
Figure 4.13: Ex-situ XRD (a) and SEM (b-g) images of all electrodes after cyclic stability
test. .................................................................................................................................... 44
Figure 4.14: (a) GCD curves, (b) Specific capacitance and current density, relation (c) EIS
of full cell, (d) CV profile, (e) Relation between capacitance and scan rate, (f) Ragone plot.
........................................................................................................................................... 46

vii
LIST OF SYMBOLS, ABBREVIATIONS AND ACRONYMS

TMC Transition Metal Chalcogenides

TMO Transition Metal Oxide

SC Supercapacitor

EDLC Electrical Double Layer Capacitor

FTIR Fourier Transform Infrared Spectroscopy

XRD X-ray Diffraction

SEM Scanning Electron Microscopy

XPS X-ray Photon Spectroscopy

TGA Thermogravimetric Analysis

BET Brunauer-Emmett-Teller

CV Cyclic Voltammetry

EIS Electrochemical Impedance Spectroscopy

GCD Galvanostatic Charge Discharge

ECSA Electrochemical Active Surface Area

EDS Energy Dispersive Spectroscopy

BJH Barrett-Joyner-Halenda

viii
ABSTRACT

TMCs have established themselves as potential electrodes for pseudocapacitors because of

their superior redox activity, short transport pathways, varied morphological, better

cyclability and numerous electrochemical active sites. FTIR, XRD, XPS, SEM, EDS, BET

and TGA were used for the characterization of prepared materials. A ternary composite

(TiO2/MOS2/SnS) electrode material was produced using a hydrothermal method. This

electrode at 1 A/g showed a high capacitance of 1519 F/g. The material maintained

impressive cyclic reliability, retaining 92.77% capacity after 10,000 (GCD) cycles. It also

displayed at 10 A/g, a coulombic efficiency of 90.4%. The energy density of 13.18 Wh/kg

and a power density of 125 W/kg were calculated. Furthermore, a full cell configuration

(two-electrode system) was constructed, achieving a capacitance of 395 F/g at 0.5 A/g. The

full cell's energy density was recorded at 79 Wh/kg, and 300 W/kg power output. This

remarkable electrochemical activity and pseudocapacitive behavior is due to synergy of its

components.

ix
CHAPTER 1: INTRODUCTION

The rising population and rapid industrialization have led to a significant surge in energy

consumption and contribute to various environmental challenges such as climate change,

pollution, and global warming [1]. Non-renewable resources posses’ threat to the

environment and hence a shift towards renewable energy is inevitable. Renewable energy

emerged as a viable solution to address energy issues and environmental problems.

However, the short-term availability of these renewable resources calls for the

development of energy storage solutions to create a balance between the demand and

supply of energy [2]. These include batteries, supercapacitors, and fuel cells. Each has its

own advantages regarding energy capability and power output and cyclic stability. Figure

1.1 compares the power density with the energy density of several energy storage devices.

Figure 1.1: Energy capability and power output [3].

1.1 Supercapacitors

1
Supercapacitors (SC) are the devices that store charge through the phenomenon of

electrostatic attraction (non-faradaic) and oxidation/reduction reactions (Faradaic) [4, 5].

Some of the properties of supercapacitors are as follows:

1. High power output capability

2. Appropriate energy storage capacity

3. Rapid charging and discharging rates

4. Extended Life cycle

5. Broad range of operating temperature

Therefore, supercapacitors find their use in applications that demand high power outputs,

including regenerative braking systems, wind turbines, and electric vehicles (EVs)

Supercapacitors have three primary categories according to the mechanism of charge

storage:

1. Electrical double layer capacitor

2. Pseudocapacitors

3. Hybrid capacitor

1.1.1 Electrical double layer capacitors (EDLCs):

EDLCs (Electric Double-Layer Capacitors) accumulate charge through electrostatic

attraction, creating a double layer. All carbon-based electrode materials show EDLC

behavior. This encompasses materials such as AC, CNTs and mesoporous and microporous

carbon [6]. There are no faradaic reactions involved and hence CV curves are rectangular

2
shaped while GCD profiles are also symmetric. The charge storage mechanism of

EDLC is illustrated in Fig. 1.2.

Figure 1.2: The charge storage mechanism of EDLC [7].

1.1.2 Pseudocapacitors

Pseudocapacitors primarily store charge through reversible redox reactions. The charge is

mostly stored in three ways: (a) adsorption, (b) redox reactions and (c) intercalation

mechanisms. In adsorption as shown in Figure 1.3, metal ions are adsorbed on electrode

surface. In Faradaic processes, the movement of charge is facilitated by reversible redox

(reduction-oxidation). In intercalation pseudocapacitors, ions are integrated into the layers

of the electrode material [7].

3
Figure 1.3: Pseudocapacitors exhibiting adsorption and faradaic reactions [7].

1.1.3 Hybrid Capacitors

They encompass both charge storage mechanisms of both Faradaic (redox reactions) and

non-Faradaic processes (adsorption). These devices merge the energy characteristic of

EDLCs with the high-power outpput found in pseudocapacitors. In recent years, scientists

have extended their research on developing hybrid supercapacitors. Hybrid supercapacitors

are further divided into battery type and composite type. Figure 1.4 illustrates the various

types of supercapacitors

4
Figure 1.4: Types of supercapacitors.

1.2 Supercapacitors based on Transition Metal Chalcogenides

Transition Metal Chalcogenides (TMCs) represent a category of compounds with the

general formulas MX and MX2, comprising transition metals and chalcogen elements.

They have established themselves as potential electrode materials for supercapacitors.

TMCs have high surface area, excellent conductivity, superior redox activity, rich in active

site and short diffusion pathways due to their layered structure [8, 9].

Apart from their vast applications in energy storage systems, they suffer from certain

limitations as well which are as follows:

1.2.1 Low conductivity

Transition Metal Chalcogenides (TMCs) generally exhibit lower conductivity including

materials like CNTs, transition metal oxides, and CP. Their conductivity can be improved

5
by developing novel heterostructures by incorporating carbon materials and conducting

polymers into them. This can combine the properties of individual components and will

enhance the performance [10].

1.2.2 Poor cycle stability

Another drawback that limits the application of TMCs is their poor cycle stability. After

long cycling, they undergo volume expansion affecting the crystal structure and hence the

performance of a supercapacitor. This can be reduced by incorporating materials that can

withstand their volume alterations such as EDLC materials and other metal-based

compounds [11].

1.2.3 Limited surface area

Compared to the EDLC and pseudocapacitive electrode materials, TMCs generally have

low surface area with limited active sites which limits its electrochemical activity. The

surfaces area of transition metal chalcogenides (TMCs) can be enhanced through

nanoengineering, which involves forming diverse morphological structures such as

nanorods, nanofibers, nanoflowers, and core-shell structures [12].

1.2.4 Aggregation and restacking issues

TMCs have layered structures held by Van der Wal forces and they can suffer from

aggregation or restacking. This reduces the effective surface area which hinders the

material’s electrochemical performance. Hence, we modify TMCs to enable them to

withstand these changes [13].

6
1.3 Modified transition metal chalcogenides

To overcome the inherent limitations of transition metal chalcogenides, several

modification techniques have been employed to enhance their electrochemical

performance and active surface area of materials for supercapacitor applications. The

modification strategies are demonstrated in Figure 1.5. Some of the modification strategies

are as follows:

1.3.1 Carbon additives

Carbon based materials have good electrical conductivity, easy fabrication, eco-friendly,

porosity and low toxicity [14]. These include graphene, activated carbon and CNTs.

Adding carbon-based materials to the TMCs, we can overcome not only improve the

conductivity but also increase the availability of active sites, provide structural integrity,

improve cycle stability, and contribute to the superior electrochemical performance [15].

1.3.2 Conductive polymers

Conducting polymers are organic compounds having conducting properties like metals and

semiconductors. They are low cost, environmentally friendly and provide mechanical

strength. Along with the high conductivity they offer mechanical flexibility to the

materials. Polyaniline, polypyrrole, poly(p-phenylenevinylene) and

polyethylenedioxythiophene are some commonly used conductive polymers [16]. By

fabricating composites with conducting polymers, we can improve the cycle stability,

structural integrity, and conductivity of the TMCs [17].

1.3.3 Heterostructure formation

7
1.3.3.1 TMCs/Metals

To overcome the limitations of TMCs, various metal nanostructures can be incorporated in

them such as graphene, conducting polymers and TMOs resulting in the superior

electrochemical activity, good cyclability, high active surface area and various oxidation

states.

1.3.3.2 TMCs/TMOs

TMOs have established themselves as promising materials for supercapacitor application.

They offer exceptional properties such as good cyclability, low cost, abundance,

environment friendliness and catalytic behavior [18]. Some TMOs such as Fe3O4, Mn3O4,

RuO2, Co3O4, NiO and TiO2 have shown excellent capacitive behavior in recent years. On

developing heterostructures with TMCs, the synergy ascension in electrochemical

performance can be achieved due to the combination of TMOs and TMCs and their

inherent limitations can be inhibited by their coordination [19, 20].

1.3.3.3 TMCs/TMCs

Designing heterostructures by integrating metal different TMCs together is another

modified strategy to obtain superior electrochemical properties by combining the

individual properties of each TMCs [8]. The synergy combination helps to limit the

restacking and volume issues of them and thus improves the overall performance. The

diverse morphologies such as core-shell, flower shape, etc. can because of a result of their

combination. Transition meta sulfides are better performers than their carbide, nitride and

phosphides.

8
1.3.3.4 Heteroatom doping

Another way to modify is by doping heteroatoms for instance nitrogen, oxygen, boron, and

phosphorus etc. Heteroatom doping increases the active sites in the materials.

Figure 1.5: An illustration of the modification strategies of TMCs for superior

electrochemical performance [21].

1.4 Objectives

➢ To develop a facile synthesis method for the high surface TiO2/MoS2/SnS ternary
composite.
➢ To investigate the synergy of combining TiO2, MoS2, and SnS in the composite and
how these effects contribute to enhance capacitance.
➢ To analyze the cycle stability of the electrode material.
➢ Assembly of a two-electrode cell to achieve optimal device performance.

9
CHAPTER 2: LITERATURE REVIEW

Iqbal et al. synthesized MoS2@TiO2 as a advanced supercapacitor electrode material

owing to outstanding properties of transition metal sulfides, for instance their high surface

area and excellent conductivity. Hydrothermal technique was opted for the synthesis of the

composite with different concentrations of TiO2. They also fabricated symmetric

supercapacitor device and whereas, TiO2 nanorods act as substrate. The electrochemical

performance was evaluated for all electrodes. Among the engineered structures, MoS2 with

a 15% TiO2 composition exhibited 210 F/g capacitance. Impressively, this material

retained 98% of its capacitance after 2000 Galvanostatic Charge-Discharge (GCD) cycles

at room temperature, demonstrating excellent cycle stability and durability as an electrode

material. The electrochemical behavior improved at elevated temperatures by a factor of

200% at 60 °C. EIS analysis showed minimum charge transfer resistance owing to enhance

supercapacitor performance. Therefore, the binary composite MoS2@ 15%TiO2

demonstrated remarkable performance over broad temperature range (25–60 °C) and can

be used as energy storage device [22].

Lathe et al. prepared TiO2@MoS2 heterostructure via facile one-step hydrothermal

process. TiO2@MoS2 achieved capacitance of 337 F/g. Additionally, the photocatalytic

behavior of the TiO2@MoS2 heterostructure was also studied and results showed that it

exhibited the showed the photoreduction of Cr6+ to Cr3+ with formic acid under direct

sunlight in 2 h. Hence, the composite has applications as supercapacitor as well as a

photocatalyst [23].

10
Chanda et al. synthesized hierarchical structure consisting of MoS2 nanoflakes supported

on TiO2 nanospheres. In the synthesis process using the two-step hydrothermal method,

TiO2 serves as the substrate onto which MoS2 nanoflakes are grown, forming the composite

material. The electrochemical studies highlighted that the composite exhibited a

remarkable capacitance value which is over 30% greater than that of the pristine spherical

shaped TiO2. This better activity can be attributed to the large surface area and improved

morphology. [24].

Sarkar et al. reported free standing nanoflakes on molybdenum foil. Hydrothermal process

was used to synthesize the composite, MoS2/r-GO and asymmetric supercapacitor device

was tested in Na2SO4 as an electrolyte. The unique morphology and layered structure

facilitate the Na+ intercalation/deintercalation and lead to higher capacitance, greater cycle

stability, and capacitive retention after 20,000 cycles. [25].

Dar et al. reported that tin sulfide nanoparticles were prepared in three solvents: ethanol,

ethylene glycol, and oleyl amine using the solvothermal technique. The process involved

amalgamating SnS nanoparticles with a mixture of polyvinyl pyrrolidone and one of the

solvents. For identification purposes, ethanol was referred to as AR1, ethylene glycol as

AR2, and oleyl amine as AR3. XRD, SEM, TEM, FTIR, UV-DRs and electrochemical

testing was performed for all three samples. XRD reveals the orthorhombic phase of SnS

while SEM indicates the morphological changes. Electrochemical testing showed that the

AR1 electrode, prepared with ethanol, exhibited the 419 F/g capacitance compared to those

prepared with AR2 (ethylene glycol) and AR3 (oleyl amine). The superior capacitance of

AR1 is attributed to its better morphology and improved surface area [26].

11
Chauhan et al. revealed the production of SnS nanorods without the use of surfactants,

employing a solvothermal technique. This method came out to be low cost and

environment friendly for studying its supercapacitor applications. In terms of

electrochemical evaluation, the device was assembled without the use of any binder,

ensuring a pure assessment of the material's capabilities. The material demonstrated an

impressive capacitance reaching up to 70 F/g [27].

Li et al. revealed a unique method of synthesizing carbon coated SnS using ball milling

and then heating it with PVA. To investigate the composite's crystal structure and surface

morphology, the electrode material underwent analysis using XRD and SEM. The

electrochemical testing was done and compared with the pristine sample, for its application

in supercapacitor. The carbon coated SnS exhibited an impressive capacitance of 28.47

F/g, surpassing the performance of uncoated, pristine SnS [28].

Huang et al. adopted a novel synthesis route using L-cysteine assisted MoS2/graphene

composite. The XRD method was employed to examine the layered architecture of the

MoS2-graphene composite, elucidating its structural properties. Concurrently, the SEM

analysis provided a detailed view of the composite's three-dimensional spherical

morphology, offering a deeper understanding of its physical form. The findings showcased

a peak capacitance of 243 F/g at 1 A/g and maintaining 92.3% capacitance for 1000 charge-

discharge cycles at the same current density. The superior cycling stability and high

specific capacitance are credited to the material's unique morphology and extensive active

surface area. These characteristics not only prevent volume expansion during charge-

discharge cycles but also enhance the ion transport from the electrolyte solution to the

12
electrode material. Given these exceptional performance metrics, the material emerges as

a promising candidate for use as an electrode in supercapacitor applications. [29].

The core-shell heterostructures were developed by Lu et al. as effective electrodes for

supercapacitors. He synthesized 3D CNF/TiO2@MoS2 hierarchical nanostructures by first

preparing carbon nanofibers by electrospinning. The TiO2@MoS2 grow via hydrothermal

process on these nanofibers because they act as a substrate. The electrochemical behavior

was inspected for the electrode and the findings highlighted the exceptional cycle stability

and high capacitance of the CNFs/TiO2@MoS2, significantly surpassing the performance

of the pristine CNFs/TiO2. A capacitance of 510 F/g at 0.5 A/g of the composite showcased

its superior energy storage capabilities. The combination of carbon nanofibers (CNFs) with

TiO2 and MoS2 enhances the material's conductivity, mechanical strength and a high

surface area. These characteristics are crucial for achieving high capacitance and cycle

stability in energy storage devices. [30].

Li et al. studied net like SnS/carbon nanostructure. The nanostructures were synthesized

by using the sol-gel method, a technique known for its ability to produce uniform and

highly pure materials. In this process, SnS nanoparticles were mixed with a resorcinol-

formaldehyde sol, a type of organic polymer precursor. The solution then undergoes a high-

temperature treatment at 650 °C. The SnS/carbon electrode was examined using XRD,

SEM, and HRTEM. It showed a high capacitance of 36.16 F/g and great cyclability. Its

high performance is due to its net-like structure and high conductivity. [31].

Rajeswari et al. fabricated MoS2@TiO2 composite in a 1:1 ratio via hydrothermal route.

The contents were then subjected to heat treatment at 500 ◦C. The physiochemical

13
properties were analyzed and compared before and after the heat treatment. The

MoS2@TiO2 displayed 325 F/g capacitance at 2 A/g. Additionally, a supercapacitor device

(ASC) was also developed. The remarkable electrochemical results of the MoS2@TiO2

holds potential for highly effective supercapacitor application [32].

Zhou et al. reported 1T-MoS2@TiO2/Ti by confining 1T-MoS2 into TiO2 nanotube

through simple hydrothermal technique. Initially 1T-MoS2 nanosheets were

hydrothermally prepared but due to its unstable phase can easily be converted to 2H-MoS2

which is more stable phase and decrease its electrochemical properties. In terms of energy

storage application, 1T-MoS2 has the highest conductivity than other 2H and 3R. Herein,

1T-MoS2@TiO2/Ti was synthesized to further enhance its performance. The findings

demonstrated a capacitance of 428.1 F/g at a current of 0.2 A/g and 97% retention after

10,000 cycles [33].

Ravuri et al. established a series of SnS/graphene with varying concentrations of graphene

using wet chemical method. All the composites were analyzed by XRD to study their

crystal structure, SEM to analyze the morphological alterations and XPS to evaluate the

chemical states and surface properties. The material exhibits a capacitance of 984 F/g.

Additionally, SnS/G demonstrated excellent cyclic stability [34].

Vinodhini et al. fabricated hybrid electrodes. CNTs and TiO2 ware incorporated into MoS2

to form hybrid MoS2/TiO2/CNTs nanocomposite. The material was analyzed by Raman,

XRD, XPS, SEM and TEM. Hybrid electrode material undergoes electrochemical testing

in 1M aqueous solution of KOH. MoS2/TiO2/CNTs demonstrated good capacitive

performance with a 1252.57 mA h/g specific capacitance at 20 mV/s greater than the

14
MoS2/CNTs and TiO2/CNTs. EIS plot was circuit fitted to find Rct values of

1.7 Ω.cm2 which is lower than the corresponding electrodes which confirms the superior

capacitive performance of the material [35].

Zhu et al. fabricated MoS2/TiO2 nanocomposite with varying concentrations of TiO2 so

that the capacitance and stability of MoS2 can be improved. The addition of TiO2 helps to

limit the volume or restacking issues of MoS2. Electrochemical analysis showed that the

top-performing electrode achieved a capacitance of 1172.9 mAh/g at 0.5 A/g. It maintained

a capacitive retention of 694 mAh/g at 2 A/g over 100 cycles. Results indicate enhanced

capacitive behavior due to the synergistic properties of the components constituting the

nanocomposite [36].

15
CHAPTER 3: EXPERIMENTAL SECTION

3.1 Chemicals

All chemicals and reagents were utilized without any processing for any of the chemicals.

Ti(C4H9O)4 (Titanium tetra-n-butoxide, 97 % pure), C5H8O2 (Acetylacetone),

Na2MoO4.2H2O (Sodium Molybdate Dihydrate, 98% pure), SnCl2.2H2O (Tin Chloride

Dihydrate, 98% pure), HCl (hydrochloric acid, 37 %) and SC(NH2)2 (Thiourea, 99% pure)

were used in the synthesis process. Super P, polyvinylidene difluoride (PVDF) and NMP

(N-methyl-2-pyrrolidone), were used for preparing slurry for electrodes. The deionized

(DI) water, Isopropyl alcohol (IPA) and ethanol were used as solvents in the synthesis

process.

3.2 Experimental

3.2.1 Synthesis of TiO2

The synthesis of TiO2 nanoparticles commenced with a hydrothermal approach. Initially,

10 mL of acetylacetone was combined with 40 mL of isopropyl alcohol, ensuring constant

agitation. Subsequently, 3 mL of titanium butoxide was introduced into the mixture, stirred

vigorously for 10 minutes, before putting in autoclave. After heating for 5 h at 190 °C, the

contents were obtained and washed multiple times with isopropyl alcohol through

centrifugation. The precipitates were vacuum dried overnight.

16
Figure 3.1: Synthesis Scheme of TiO2

3.2.2 TiO2/MoS2 binary composite (TM)

The TiO2/MoS2 composite was created by dissolving sodium molybdate and thiourea in a

1:4 mmol ratio in 40 mL of deionized (DI) water, with the mixture subjected to intense

stirring. 50 mg of as prepared TiO2 powder was dispersed ultrasonically in DI water. The

prepared suspension was introduced to the solution and agitated for 1 h. Subsequently, the

resultant solution underwent heated for 26 h at 190 °C. After the heating process, the black

precipitates that formed were collected. These precipitates were then subjected to multiple

washes, initially with water followed by ethanol, to ensure thorough cleaning. Finally, the

washed contents were dried. Pure MoS2 was also prepared without the addition of TiO2.

17
Figure 3.2: Synthesis Scheme of TiO2/MoS2

3.2.3 TiO2/SnS binary composite (TS)

For the synthesis of TiO2/SnS, we used the tin chloride dihydrate as a precursor of Sn.

SnCl2.2H2O and thiourea in 1:1.5 mmol ratio were dissolved in DI water (40 mL) under

continuous stirring. 50 mg of as prepared TiO2 powder, dispersed ultrasonically in a

separate beaker. The final solution mixture was enhanced by adding the specified reactants

and then stirred for 1 h and heated for 16 h at 190 °C. After the heating process, the contents

were collected and allowed to cool. The cooled mixture was washed 2-3 times, initially

with deionized (DI) water and subsequently with ethanol, to ensure thorough purification.

The greyish precipitates obtained from this process were then dried to remove any residual

18
moisture, resulting in the final product. Pure SnS was also synthesized for comparison

without the addition of TiO2.

Figure 3.3: Synthesis Scheme of TiO2/SnS

3.2.4 TiO2/MoS2/SnS (TMS)

The ternary composite was synthesized using hydrothermal technique. First separate

solutions were prepared by mixing each precursor of Sn, Mo, and S in DI water.

Na2MoO4.2H2O, SnCl2.2H2O and thiourea in a 3:6:12 were dissolved separately in DI

water. After preparing the individual solutions, they were combined to create a final

solution, which was then stirred continuously for 1 hour. To modify the pH of this solution,

a dilute HCl solution was added drop by drop until the desired pH level was achieved.

19
Following the pH adjustment, the final solution was put to an autoclave and subjected to

heating at 190 °C for 26 hours. Once the heating process was complete, the contents were

allowed to cool. The next step involved washing the cooled contents through

centrifugation, a process designed to separate the precipitates from the supernatant

efficiently. After the centrifugation, the precipitates were collected and transferred to a

drying oven, where they were dried at 60 °C for 12 hours to ensure all moisture was

removed, resulting in the final dried product.

Figure 3.4: Synthesis Scheme of TiO2/MoS2/SnS

3.3 Material characterization

20
FTIR, a technique to determine the functional groups in the synthesized samples was

performed through ATR-alpha-FTIR instrument in 400-4000 cm-1 range. XRD (X-ray

Diffraction) in the range of 2θ = 5° to 80° was conducted to study the crystallinity and

phase structure by using XRD, Bruker D8 Advance, Germany Instrument. SEM was

utilized to investigate the morphological properties using the instrument SEM, Tescan

Vega 3. To examine the electrodes' surface chemical composition, X-ray Photoelectron

Spectroscopy (XPS) was performed. It was done by using XPS, PHI 5000 VersaProbe

instrument. BET and BJH pore size distribution profiles were conducted through

Quantachrome, Novawin equipment for determining the surface area of electrodes. The

weight loss with temperature and thermal stability were evaluated by using TGA SDT650

instrument.

3.4 Preparation of the electrode material

The working electrodes were prepared by applying a thin slurry layer onto nickel (Ni)

foam, which served as the substrate. The dimensions of the Ni foam used were 1 × 2 cm².

This method utilized Ni foam as a base to create the electrodes, ensuring a suitable surface

for the deposition process. First, nickel foam was washed through ultrasonication with 1M

HCl for 15 min, then with DI water and in the end with ethanol. The washed Ni foam was

then dried. The binder solution was prepared by dissolving it in NMP under constant

stirring at 45 °C for 24 h. Active material, super p and PVDF in 7:2:1 ratio was mixed in

motor and pestle to prepare the slurry. A thin slurry layer was applied to a 1 cm² area of

nickel (Ni) foam. To assess the amount of slurry deposited, the weight of the Ni foam was

recorded both before and after the coating process. This measurement helps in determining

the precise mass of the slurry added to the foam. Mass loading was calculated by

21
subtracting the bare weight from the final weight of Ni foam. The mass loading was found

to be less than 2 mg. The electrodes that were prepared were left to dry, allowing the

process to take place overnight.

22
CHAPTER 4: RESULTS AND DISCUSSIONS

4.1 FTIR analysis

FTIR is a technique used to determine the functional groups present in a material. The

spectrum shows absorption or emission bands obtained when the material absorbs or emits

infrared radiation, and that band indicates the presence of a certain functional group.

The FTIR spectra in 400-4000 cm-1 range is depicted in Figure 4.1. The absorption bands

observed at 3410 cm-1 OH stretching and 1629 cm-1 for OH bending vibrations. These

bands occur due to absorbed water molecules within the sample [37]. These two absorption

bands were observed in all the samples.

In the FTIR spectra of TM, the absorption band at 472 and 589 cm-1 correspond to S−S and

Ti−O−Ti stretching vibrations. [38, 39]. The observed multiple absorption bands within

the range of 700-1150 cm-1, specifically at 893, 936, 1027, and 1109 cm-1, are due to the

vibrations of sulfur groups [40]. Additionally, the absorption band at 1403 cm-1 is attributed

to S−Mo−S linkage, suggesting the incorporation of molybdenum alongside sulfur in the

structure [41]. In case of TS, the band seen at 640 cm-1 is indicative of Ti−O−Ti bridge

stretching vibrations, pointing to the presence of titanium-oxygen-titanium linkages within

the material's structure [42]. The 1030 cm-1 band and 1260 cm-1 band are due to the sulfur

groups [43, 44]. In the FTIR spectra of the ternary composite TMS, the absorption bands

appear at 532, 657, 896, 942, 1035, 1132, and 1411 cm-1. The band 532 cm-1 is attributed

to the stretching vibrations of disulfide (S-S) bonds. Additionally, the one at 657 cm-1 is

ascribed to the Ti–O–Ti bond. These spectral features highlight the complex chemical

23
structure and interactions within the material. Sulfur groups have stretching vibrations

present between the 800-1150 cm-1 range. The S−Mo−S bond stretching of MoS2 in TMS

results in band at 1411 cm-1.

Figure 4.1: FTIR of TM, TS and TMS composites.

4.2 X-ray diffraction (XRD) analysis

XRD patterns of TM, TS and TMS are shown in Figure 4.2. Since TiO2 is a component of

all three synthesized materials, its characteristic major peaks are observable in the XRD

patterns of TM, TS, and TMS.

24
XRD pattern of TM exhibits peaks of both TiO2 and MoS2. The peaks at 2θ = 25.23°,

37.12°, 54.13° and 62.65° are attributed to the (101), (004), (105), and (204) planes of TiO2

anatase phase, respectively (JCPDS card no. 073-1764, space group: I41/amd) [45]. The

broad peaks at 12.3°, 28.7, ° 32.4°, 35.28°, 37.79°, 47.9° and 57.53° correspond to (003),

(006), (101), (012), (104), (107) and (110) lattice planes in the rhombohedral phase MoS 2

that matches with the standard (JCPDS card no. 077-0341, space group: R3m) [46, 47].

The major peak of MoS2 at 12.3° can be seen to be shifted to somewhat lower 2θ value

compared to the standard 14.4° which shows its expanded interlayer spacing [48]. In TS,

the diffraction peaks of TiO2 are present at 25.26°, 37.86°, 47.89° and 54.12° (JCPDS card

no. 073-1764, space group: I41/amd) [49]. The sharp peaks at 22.05°, 26.02°, 27.43°,

30.47°, 31.46°, 31.86°, 38.98°, 42.59°, 44.69°, 45.51°, 48.5°, 51.24°, 53.07°, 53.42°,

56.57°, 59.39°, 64.14°, 66.51, ° 68.74°, 72.85° and 75.17° exhibited the (110), (120), (021),

(101), (111), (040), (140), (210), (141), (002), (211), (151), (122), (160), (042), (250),

(251), (171), (232), (162) and (113) crystal planes of herzenbergite (SnS) orthorhombic

phase, respectively (JCPDS card no. 033-1375, space group: Pbnm) [50]. The ternary

composite TMS contains all the individual major peaks of TiO2, MoS2 and SnS. The peaks

of TiO2 are at 2θ = 25.53°, 37.7° and 47.95° and the peaks at 14.9°, 33.78° and 48.2° are

ascribed to MoS2 [51]. The peaks at 2θ = 26.44°, 28.22°, 31.44°, 33.78°, 39.11°, 41.82°,

44.7°, 51.55°, 54.67° can be indexed to the orthorhombic SnS. The presence of these peaks

in TMS without any shifting indicates that the ternary composite was successfully

synthesized.

25
Figure 4.2: XRD pattern of TM, TS and TMS composites.

4.3 X-ray photoelectron spectroscopy (XPS) analysis

XPS serves to explore surface chemistry, detailing the composition and states of materials.

It further delineates the chemical states of all elements found within a sample. Figure 4.3

displays the survey spectra of all three synthesized materials, indicating the presence of Ti,

Mo, Sn, O and S in them. The presence of F, C and Cu is because XPS was conducted on

the electrode instead of powder sample. The electrode was developed by depositing slurry

made up of material, super p and PVDF in a 70:20:10 ratio.

26
Figure 4.3: Survey Spectra of TM, TS and TMS composites.

The HR-XPS of S 2p is present in Figure 4.4(a). In TM, the peaks observed at 161.8 for S

2p3/2 and 163.0 eV for S 2p1/2, respectively. For TS, the distinctive peaks at 162.0 eV for S

2p3/2 and at 163.3 eV for S 2p1/2 are of sulfur in -2 oxidation state, respectively [48].

Another peak at approx. 165 eV is due to the presence of bridging S22-. In case of TMS,

the peaks at 161.9 for S 2p3/2 and 163.1 eV for S 2p1/2 which confirms S2- present in the

ternary composite [51]. A doublet at 164.3 eV and 165.4 eV is due to the bridging S22- or

27
apical S2- [52]. In all three materials, a peak at 169 eV correspond to the S−O bond due to

oxidation of metal sulfide [47]. The XPS spectra of O 1s is illustrated in Figure 4.4(b). In

TM, the peaks observed at 530.3, 531.8 and 533 eV are for Ti−O−Ti bond, oxygen

vacancies (Ovs) neighboring Ti3+ and surface adsorbed water, respectively. For TS, the O

1s peaks are present at the binding energies of 530.8, 532.5 and 533.4 eV while for TMS,

the peak corresponding to the oxygen of TiO2 is present at 530.7 eV. The other two peaks

at 531.9 and 532.7eV are due to the Ovs around Ti3+ and OH groups on the surface,

respectively. [45]. The HR-XPS spectra of Ti 2p is shown in Figure 4.4(c). For the ternary

composite, the characteristic peaks of Ti at 459.2 for Ti 2p3/2 and 461.1 eV for Ti 2p1/2

which confirms Ti in +4 oxidation state [53]. Another minor peak at 464.9 eV is due to the

interaction of Ti with sulfur (Ti-S) [54]. Figure 4.4(d) shows XPS spectra of Mo 3d. For

TMS, the peaks at 229.3 for Mo4+ 3d5/2 and 232.9 eV for Mo4+ 3d3/2 confirming the

presence of molybdenum in +4 oxidation state in MoS2 [55]. An additional peak at binding

energy of 236.2 eV refer to presence of Mo6+ [56]. This peak may arise due to the surface

oxidization of MoS2 [46, 47]. The peak of S 2s orbital can be observed at 226.6 eV [47,

57]. The Sn 3d XPS spectra is illustrated in Figure 4.4(d). In TMS, the peaks at 487.2 eV

and 495.6 eV for Sn2+ 3d3/2 and Sn2+ 3d3/2 refer to Sn−S bond [50]. These shift in the peaks

to lower binding energies from the corresponding peaks in the binary composite TS (487.3

and 495.7 eV) indicate the optimization of electronic structure in the ternary composite

[58].

28
Figure 4.4: (a) S, (b) O, (c) Ti, (d) Mo, and (e) Sn, XPS spectra respectively.

4.4 SEM and EDS analysis

The structural characteristics of the three materials were further analyzed through SEM to

gain insights into their morphological features. SEM micrographs of TM, TS and TMS are

shown in Figure 4.5. The SEM images show irregular and agglomerated heterostructures

at high magnifications. At low magnification, these heterostructures appear to be somewhat

flower-like or disc like secondary particles. Small primary particles aggregate to form the

secondary particles. The small porous particles can be seen dispersed on the surface of the

large secondary ones [48, 50]. The porous surface observed in the materials suggests an

increased number of active sites. This improvement can lead to better performance where

efficient ion exchange is crucial [59]

29
Figure 4.5: (a, d) TM, (b, e) TS and (c, f) TMS, SEM images.

EDS spectra of the ternary composite is illustrated in Fig 4.6. Elemental mapping confirms

that O, Ti, Mo, Sn and S are present without any other impurity element which indicates

the successful synthesized of ternary composite by hydrothermal technique.

Figure 4.6: EDS Spectra and elemental of TMS composite.

4.5 Thermogravimetric analysis (TGA)

30
TGA analysis was done to study the mass loss with the increase in temperature. The

stability of all materials was investigated by heating them in an inert atmosphere from 20

°C to 900 °C. TGA of all three samples is shown in

Figure 4.7. The initial weight loss observed below 200 °C is attributed to the loss of

adsorbed water on the material. Following this, a second weight loss in 200-350 °C, which

can be ascribed to the thermal decomposition of metal sulfates present in the sample. This

two-step weight loss process is indicative of the material's thermal stability and the nature

of its components. The final loss from 350 °C to 900 °C indicates the degradation of MoS2

along with the phase change of TiO2 from anatase to rutile. [37]. For TS, the composite is

quite stable for a wide range of temperatures and a sudden weight loss starts from 700 °C

which demonstrates its excellent thermal stability. By integrating SnS, TiO2 and MoS2 in

the ternary composite, we attain good thermal stability for our ternary composite.

Figure 4.7: TGA of TM, TS and TMS composites.

31
4.6 BET analysis

All synthesized materials were analyzed using the BET and BJH pore techniques. Figure

4.8(a) shows the isotherms for all the samples. The adsorption isotherms presented for all

the samples indicate that they adhere to type IV isotherm curves. According to the IUPAC

classification, hysteresis loop of H3 type was observed. This suggests the presence of

mesopores that are accessible for adsorption and capillary condensation. BET surface area

was 81.593 m2/g for TMS greater than 66.017 m2/g and 8.387 m2/g for TS and TM

composites, respectively. BJH pore size distribution (PSD) as seen in Figure 4.8(b) is

attributed to the presence of mesopores and micropores in the samples. Table 4.1 display

surface area and pore data of all prepared samples. The improved electrochemical behavior

of TMS can be linked to its increased surface area and pore volume, which enhances the

charge transfer across the material’s surface.

Table 4.1: Analysis of surface area and BJH pore data of TM, TS, and TMS composites.

Sample Electrode Surface Volume of pore Size of pore (nm)


area (m2/g) (cc/g)

TM 8.387 0.036 1.5906

TS 66.017 0.178 2.6824

TMS 81.593 0.186 1.8108

32
Figure 4.8: (a) BET sorption isotherms, (b) BJH pore data of TM, TS and TMS
composites.

4.7 Cyclic Voltammetry (CV) analysis

Cyclic voltammetry (CV) was used to determine the electrochemical performance of all

samples. Cyclic voltammetry (CV) was conducted within a voltage range of 0.35–0.6V

(∆V = 0.25 V) using a KOH solution (3 M). The CV patterns for all the tested samples,

captured at varying scan rates as illustrated in Figure 4.9(a-c). Each CV pattern showcases

redox peaks, highlighting the pseudocapacitive characteristics of the electrode materials.

These redox peaks are more pronounced at low scan rates, whereas CV profiles tend to

adopt a more rectangular shape at high scan rates, indicating a relationship between the

curve's area and the scan rate. At 5 mV/s, a comparison of all electrodes is shown in Figure

4.9(d), revealing that the TMS electrode exhibits the largest area under the curve,

signifying enhanced electrochemical activity. Following equation (1) was used for

calculating the specific capacitance of all electrodes.

33
𝐴
𝐶𝑠 = (1)
2𝑚𝑘(∆𝑉)

Table 4.2 presents the specific capacitance values for all electrodes, determined using the

previously mentioned formula. It is observed that the capacitance exhibits an inverse

relationship with the scan rate. At high scans, the capacitance decreases due to fast charge

discharge reactions, and it increases at low sweep rates due to ions having more time to

intercalate through the electrode material.

Table 4.2: Specific capacitances of TM, TS, and TMS electrodes calculated from CV.

Electrodes Specific Capacitance (F/g)

Scan Rate 5 10 20 40 60 80 100

(mV/s)

TM 469 376 267 164 118 91 71

TS 585 406 270 177 134 112 96

TMS 972 814 630 441 335 274 231

The TMS electrode at 5 mV/s has capacitance value of 972 F/g compared to the binary

composites TM and TS. This superior performance was due to the synergistic combination

of TiO2, MoS2 and SnS which ease the intercalation of ions into the layered structure.

34
Electrochemical active surface area (ECSA) of all the electrodes is calculated by using

their CV profiles. A linear graph was plotted between scan rate and ∆J which is equal to

the half of anodic and cathodic currents difference in the non-faradaic region. The slope of

this line is called Cdl as shown in Figure 4.9(f). ECSA was calculated using the following

equation (2).

𝐶𝑑𝑙
𝐸𝐶𝑆𝐴 = (2)
𝐶𝑠

where Cs indicates the capacitance of reference material. The Cs value for basic electrolyte

is 0.040 mF cm -2. Among the three electrodes, TMS has the greatest ECSA of 908 cm2 due

to its larger BET surface area and BJH pore size.

Figure 4.9: CV profiles between 5–100 mV/s of (a) TM, (b) TS, and (c) TMS electrodes.

(d) CV curves compared at 5 mV/s of all electrodes. (e) The capacitance values of

35
electrodes at all scan rates. (f) Electrochemical active surface area (ESCA) of all three

electrodes.

Dunn’s formula was used to analyze the charge storage contributions from the capacitive

and diffusion currents using CV profiles [22].

The formula is as follows:

1
𝑖 = 𝑘1 𝑣 + 𝑘2 𝑣 2 (3)

where i is current at constant V, k2ν indicates the capacitive contribution and the capacitive

current (i) is directly related to the scan rate. k1v1/2 gives the diffusion contribution [60].

Rearranging Eq. (3) to obtain eq. (4)

𝑖 1⁄
1 = 𝑘1 𝑣 2 + 𝑘2 (4)
𝑣 2

The capacitive and diffusion contribution from 5-100 mV/s for each electrode as indicated

in Figure 4.10(a-c). Capacitive behavior predominates the diffusion process at high sweep

rates because of the fast charge discharge process. TMS shows a capacitive contribution of

63% while TM exhibits 17% and TS shows 27% at a scan rate of 100 mV/s. TMS electrode

outstand the binary composites and hence increase the electrochemical activity. A

comparison of capacitive and total current at 40 mV/s is illustrated in Figure 4.10(d–f).

36
Figure 4.10: Capacitive and diffusion contribution ratios between 5-100 mV/s for (a) TM,

(b) TS, and (c) TMS electrodes. A comparison of capacitive and total current at 40 mV/s

for (d) TM, (e) TS and (f) TMS electrodes.

4.8 Galvanostatic charge-discharge (GCD) analysis

GCD is a recognized method used for measuring the capacitance of electrode materials.

Specific capacitance is computed by using equation (5):

𝐼×𝑡
𝐶= (5)
𝑚 × ∆𝑉

The GCD w conducted at different current densities for all electrodes within a voltage

range of 0.35V to 0.6V, as illustrated in Figure 4.11(a-c). All electrodes exhibit

asymmetric curves due to their faradaic nature. The discharge curves exhibit three distinct

regions, IR drop, voltage plateau and a sharp discharge curve. The GCD curve shows that

37
TMS electrode exhibits minimal IR drop (1 V) compared to TM and TS (1.7 V). The

capacitance values of all three electrodes are given in Table 4.3. Figure 4.11(d) displays a

comparison of the GCD curves at 1 A/g. At 1 A/g, the ternary composite shows the greatest

capacitance value of 1519 F/g as it takes more time to discharge compared to TM and TS

electrodes. This allows efficient ion transport during charging/discharging contributing to

it superior electrochemical properties. The TMS composite has a 90.4 % coulombic

efficiency at 10 A/g.

Table 4.3: Specific capacitances of TM, TS, and TMS at varying current densities.

Electrodes Specific Capacitance (F/g)

Current 1 2 3 5 7 10

Density

(A/g)

TM 679 642 613 561 409 348

TS 1095 902 792 663 586 526

TMS 1519 1125 1094 1047 991 740

Equations (6) and (7) is as follows:

(𝐶𝑠 ∆𝑉 2 )
𝐸𝑠 = (6)
2 × 3.6

38
3600𝐸𝑠
𝑃𝑠 = (7)
∆𝑡

Figure 4.11(f) illustrates the Ragone plot. The ternary composite TMS has the highest

energy density for a constant value of power density which makes it an excellent candidate

of high power/high energy applications. TMS has a 13.18 Wh/kg energy density while TM

has 5.89 Wh/kg and TS has 9.51 Wh/kg at 1 A/g.

The stability test was performed for 10,000 GCD cycles at 10 A/g. After completing the

10,000 cycles, the ternary composite demonstrated a retention of its capacitance value at

92.77%, as depicted in Figure 4.12(a). The capacitance drops in the first 2000 cycles,

however, it becomes constant afterwards. This exceptional retention is due to the combined

effects of TMOs and TMCs. Figure 4.12(a) inset depicts the comparative GCD curve of 1st

cycle and 10,000th cycle showing negligible changes between them which confirms the

consistent capacitive behavior. Comparison of specific capacitance values and cycle

stabilities of similar materials in the literature with our work is illustrated in Table 4.5.

39
Figure 4.11: GCD curves between 1–10 A/g of (a) TM, (b) TS, and (c) TMS electrodes.

(d) Comparative GCD curves at 1 A/g. (e) Specific capacitances derived from GCD curves

at all current densities. (f) Ragone plot for the three electrodes.

4.9 EIS analysis

Electrochemical impudence spectroscopy (EIS) analysis was conducted to investigate the

dynamics of charge kinetics and internal resistance of all three electrodes. Nyquist plot of

all the three electrodes in Figure 4.12(b) has three components. The inset of Figure 4.12(b)

displays the circuit used for fitting. The Rs and Rct calculated values are given in Table 4.4.

The solution resistance values for TM, TS and TMS are 0.9978 Ω, 0.798 Ω and 0.8481 Ω,

respectively. The Rct values are 28.08 Ω, 20.16 Ω and 3.035 Ω for TM, TS and TMS

respectively. TMS has a remarkably low Rct value of 3.035 Ω which refers to the improved

mobility and rapid movement of ions when compared to the TM and TS electrodes.

40
Table 4.4: Rs and Rct values of TM, TS and TMS electrodes.

Electrodes Rs (Ω) Rct (Ω)

TM 0.9978 28.08

TS 0.798 20.16

TMS 0.8481 3.035

41
Figure 4.12: (a) Stability test of TMS for 10,000 cycles at 10 A/g, and (b) EIS of TM, TS,

and TMS electrodes.

Table 4.5: Specific capacitances of similar electrodes as reported in literature.

Electrode Electrolyte Specific Capacitance Cyclic Stability Ref.


(F/g), x=current
density

MoS2-Graphene 1 M Na2SO4 243, x= 1 A/g 92.3% for 1000 cycles [29]


@1 A/g

CNFs/TiO2@MoS2 KOH (3 M) 510.4, x= 0.5 A/g 95.7% for 3000 cycles [30]
@3.5 A/g

SnS KOH (1 M) 201, x= 0.1 A/g 88% for 1000 cycles [61]
@1 A/g

42
MoS2 / NG KOH (6 M) 245, x= 0.25 A/g 91.3% for 1000 cycles [62]
@ 2 A/g

SnS2/ SnS /N-CNO 2 M KOH 742, x= 0.5 A/g 95% for 2000 cycles [63]
@2 A/g

TiO2/MoS2/SnS 3 M KOH 1519, x= 1 A/g 92.77% for 10,000 This


cycles @ 10 A g-1 work

4.10 Ex-situ XRD and SEM Analysis

The structural analysis of all three electrodes after stability test was conducted by using

XRD and SEM. XRD pattern presents crystal structure of all electrodes after the stability

test as shown in Figure 4.13(a). All the major individual peaks of TiO2, MoS2 and SnS can

be seen in the XRD pattern. Some peaks arise due to the nickel foam which was used as a

substrate. The peaks observed at 44.5°, 51.8°, and 76.3° are indicative of the crystalline

structure of Nickel (JCPDS card no. 004-0850, space group: Fm3m). A minor shift in the

peak positions was observed compared to the original. This shift is due to the continuous

cycling of the material at high current density. The SEM micrographs are shown in Figure

4.13(b-g) reveals agglomeration of the material and some cracks appear too because of the

mechanical degradation of the material after cycling. However, the pores can be seen in

the material which indicates that electrode has retained its structural integrity and surface

morphology.

43
Figure 4.13: Ex-situ XRD (a) and SEM (b-g) images of all electrodes after cyclic stability

test.

4.11 Full Cell Testing

Two-electrode cell was fabricated to obtain optimum device performance. The full cell was

assembled. The electrode was developed by depositing a slurry consisting of material

(TMS), super p, and PVDF in a 70:20:10 ratio of on 1 × 2 cm2 stainless-steel substrate. A

PP (polypropylene) separator dipped in 3 M KOH was placed between the electrodes.

Similarly counter electrode was fabricated by depositing slurry of activated carbon. The

device was used for 2-electrode testing. Electrochemical evaluations were conducted

within -0.6 V to 0.6 V window, and EIS was carried out across in range of 1 Hz to 105 Hz.

The galvanostatic charge discharge (GCD) curves were obtained at multiple current

densities are indicated in Figure 4.14: (a) GCD curves, (b) Specific capacitance and current

44
density, relation (c) EIS of full cell, (d) CV profile, (e) Relation between capacitance and

scan rate, (f) Ragone plot(a). The specific capacitance values are listed in Table 4.6.

Table 4.6: Specific Capacitance of full cell 0.5-10 A/g.

Electrodes Specific Capacitance (F/g)

Current 0.5 1 2 3 5 7 10
Density
(A/g)

Full Cell 395 256 101 57 12 6 3

The full device demonstrated a 395 F/g capacitance when tested at 0.5 A/g. Figure 4.14(b)

shows the capacitance value rises as the current density decreases. CV profile at various

sweep rates as in Figure 4.14(d). The Nyquist plot was circuit fitted to find Rs value of

0.9711 Ω and Rct value of 2.092 Ω as presented in Figure 4.14(c). The Ragone plot, which

is presented in Figure 4.14(f), shows that the complete cell achieves an energy of 79 Wh/kg

and 300 W/kg power output. These values were derived using equations (6) and (7).

45
Figure 4.14: (a) GCD curves, (b) Specific capacitance and current density, relation (c) EIS

of full cell, (d) CV profile, (e) Relation between capacitance and scan rate, (f) Ragone plot.

46
CHAPTER 5: CONCLUSION

In conclusion, TiO2/MoS2/SnS ternary composite was successfully synthesized via

hydrothermal technique, compared with the TiO2/MoS2 and TiO2/SnS binary composites.

The ternary composite showcased a significant capacitance of 1519 F/g at 1 A/g, along

with cyclic stability, maintaining 92.77% of its capacity for 10,000 cycles. Additionally,

the material demonstrated a coulombic efficiency of 90.4% when tested at 10 A/g. The

ternary composite has an energy capability of 13.18 Wh kg-1 and 125 W kg-1 power output.

EIS results of ternary hybrid showed minimum impedance, compared to the other binary

electrodes. 2-electrode testing was also performed for the full device. At 0.5 A g-1, it

demonstrated a capacitance value of 395 F g-1. The energy and power density of the full

cell was calculated to be 79 Wh/kg and 300 W/kg, respectively. The remarkable

electrochemical performance and pseudocapacitive behavior is due to synergistic

interaction among its components.

47
REFERENCES

[1] J. P. Dorian, H. T. Franssen, and D. R. J. E. p. Simbeck, "Global challenges in


energy," vol. 34, no. 15, pp. 1984-1991, 2006.

[2] P. Simon, Y. Gogotsi, and B. J. S. Dunn, "Where do batteries end and


supercapacitors begin?," vol. 343, no. 6176, pp. 1210-1211, 2014.

[3] S. Sagadevan et al., "Background of energy storage," in Advances in


Supercapacitor and Supercapattery: Elsevier, 2021, pp. 1-26.

[4] F. Guo, N. Gupta, X. J. S. T. Teng, and P. Solutions, "Enhancing pseudocapacitive


process for energy storage devices: analyzing the charge transport using electro-
kinetic study and numerical modeling," vol. 87, 2018.

[5] R. Barik, V. Tanwar, and P. P. Ingole, "Supercapacitors: Future direction and


challenges," in Nanostructured materials for supercapacitors: Springer, 2022, pp.
619-644.

[6] J. Libich, J. Máca, J. Vondrák, O. Čech, and M. J. J. o. e. s. Sedlaříková,


"Supercapacitors: Properties and applications," vol. 17, pp. 224-227, 2018.

[7] S. Sagadevan et al., "Fundamental electrochemical energy storage systems," in


Advances in supercapacitor and supercapattery: Elsevier, 2021, pp. 27-43.

[8] M.-R. Gao, Y.-F. Xu, J. Jiang, and S.-H. J. C. S. R. Yu, "Nanostructured metal
chalcogenides: synthesis, modification, and applications in energy conversion and
storage devices," vol. 42, no. 7, pp. 2986-3017, 2013.

[9] J. Theerthagiri et al., "Recent advances in metal chalcogenides (MX; X= S, Se)


nanostructures for electrochemical supercapacitor applications: a brief review,"
vol. 8, no. 4, p. 256, 2018.

48
[10] Y. Kim et al., "Layered transition metal dichalcogenide/carbon nanocomposites for
electrochemical energy storage and conversion applications," vol. 12, no. 16, pp.
8608-8625, 2020.

[11] J. Cherusseri, N. Choudhary, K. S. Kumar, Y. Jung, and J. J. N. H. Thomas, "Recent


trends in transition metal dichalcogenide based supercapacitor electrodes," vol. 4,
no. 4, pp. 840-858, 2019.

[12] R. Liu et al., "Fundamentals, advances and challenges of transition metal


compounds-based supercapacitors," vol. 412, p. 128611, 2021.

[13] A. Kumar et al., "Asymmetric supercapacitors with nanostructured RuS2," vol. 35,
no. 15, pp. 12671-12679, 2021.

[14] L. L. Zhang and X. J. C. s. r. Zhao, "Carbon-based materials as supercapacitor


electrodes," vol. 38, no. 9, pp. 2520-2531, 2009.

[15] L. Li, Y. Huang, and Y. J. E. Li, "Carbonaceous materials for electrochemical CO2
reduction," vol. 2, no. 1, p. 100024, 2020.

[16] A. K. Thakur et al., "Facile synthesis and electrochemical evaluation of


PANI/CNT/MoS2 ternary composite as an electrode material for high performance
supercapacitor," vol. 223, pp. 24-34, 2017.

[17] G. A. Snook, P. Kao, and A. S. J. J. o. p. s. Best, "Conducting-polymer-based


supercapacitor devices and electrodes," vol. 196, no. 1, pp. 1-12, 2011.

[18] R. Liang et al., "Transition metal oxide electrode materials for supercapacitors: a
review of recent developments," vol. 11, no. 5, p. 1248, 2021.

[19] B. Gao et al., "MnO2–TiO2/C nanocomposite arrays for high-performance


supercapacitor electrodes," vol. 584, pp. 61-65, 2015.

49
[20] A. Ramadoss and S. J. J. I. j. o. h. e. Kim, "Hierarchically structured TiO2@ MnO2
nanowall arrays as potential electrode material for high-performance
supercapacitors," vol. 39, no. 23, pp. 12201-12212, 2014.

[21] Y. Dahiya, M. Hariram, M. Kumar, A. Jain, and D. J. C. C. R. Sarkar, "Modified


transition metal chalcogenides for high performance supercapacitors: Current
trends and emerging opportunities," vol. 451, p. 214265, 2022.

[22] M. Iqbal et al., "High-performance supercapacitor based on MoS2@ TiO2


composite for wide range temperature application," vol. 883, p. 160705, 2021.

[23] A. Lathe, A. Ansari, R. Badhe, A. M. Palve, and S. S. J. A. o. Garje, "Single-step


production of a TiO2@ MoS2 heterostructure and its applications as a
supercapacitor electrode and photocatalyst for reduction of Cr (VI) to Cr (III)," vol.
6, no. 20, pp. 13008-13014, 2021.

[24] K. Chanda et al., "Hierarchical heterostructure of MoS2 flake anchored on TiO2


sphere for supercapacitor application," in AIP Conference Proceedings, 2018, vol.
1953, no. 1: AIP Publishing.

[25] D. Sarkar et al., "Expanding interlayer spacing in MoS2 for realizing an advanced
supercapacitor," vol. 4, no. 7, pp. 1602-1609, 2019.

[26] M. A. Dar, M. Y. Bhat, N. A. Mala, H. A. Rather, S. Venkatachalam, and N. J. M.


T. P. Srinivasan, "Structural, morphological and supercapacitor applications of SnS
nanomaterials prepared in three different types of solvents," vol. 66, pp. 1689-1698,
2022.

[27] H. Chauhan, M. K. Singh, S. Hashmi, and S. J. R. A. Deka, "Synthesis of surfactant-


free SnS nanorods by a solvothermal route with better electrochemical properties
towards supercapacitor applications," vol. 5, no. 22, pp. 17228-17235, 2015.

[28] Y. Li, H. Xie, and J. J. J. o. S. S. E. Wang, "Preparation and electrochemical


performances of carbon-coated nanoscale SnS for supercapacitors," vol. 15, pp.
1115-1119, 2011.
50
[29] K.-J. Huang et al., "Layered MoS2–graphene composites for supercapacitor
applications with enhanced capacitive performance," vol. 38, no. 32, pp. 14027-
14034, 2013.

[30] F. Lu, J. Wang, X. Sun, Z. J. M. Chang, and Design, "3D hierarchical carbon
nanofibers/TiO2@ MoS2 core-shell heterostructures by electrospinning,
hydrothermal and in-situ growth for flexible electrode materials," vol. 189, p.
108503, 2020.

[31] Y. Li, H. Xie, and J. J. M. L. Tu, "Nanostructured SnS/carbon composite for


supercapacitor," vol. 63, no. 21, pp. 1785-1787, 2009.

[32] M. Rajeswari et al., "Design and fabrication of high performance supercapacitor


based MoS2@ TiO2 composite electrode for wide range temperature applications,"
vol. 805, p. 139936, 2022.

[33] J. Zhou et al., "1T-MoS2 nanosheets confined among TiO2 nanotube arrays for
high performance supercapacitor," vol. 366, pp. 163-171, 2019.

[34] S. Ravuri, C. A. Pandey, R. Ramchandran, S. K. Jeon, and A. N. J. I. J. o. N. Grace,


"Wet chemical synthesis of SnS/graphene nanocomposites for high performance
supercapacitor electrodes," vol. 17, no. 01n02, p. 1760022, 2018.

[35] S. Vinodhini and J. R. J. I. J. o. E. R. Xavier, "Novel synthesis of layered


MoS2/TiO2/CNT nanocomposite as a potential electrode for high performance
supercapacitor applications," vol. 46, no. 10, pp. 14088-14104, 2022.

[36] W. Zhu, K. Liu, B. Zhang, Z. Wang, and Y. J. C. I. Wang, "Unveiling the effects
of different component ratios on the structure and electrochemical properties of
MoS2/TiO2 composites," 2024.

[37] S. Kite et al., "Nanostructured TiO2 sensitized with MoS2 nanoflowers for
enhanced photodegradation efficiency toward methyl orange, ACS Omega 6 (2021)
17071–17085," ed.

51
[38] K. Lalithambika, K. Shanmugapriya, and S. J. A. P. A. Sriram, "Photocatalytic
activity of MoS 2 nanoparticles: an experimental and DFT analysis," vol. 125, pp.
1-8, 2019.

[39] P. Praveen, G. Viruthagiri, S. Mugundan, N. J. S. A. P. A. M. Shanmugam, and B.


Spectroscopy, "Structural, optical and morphological analyses of pristine titanium
di-oxide nanoparticles–Synthesized via sol–gel route," vol. 117, pp. 622-629, 2014.

[40] G. A. Ali, M. R. Thalji, W. C. Soh, H. Algarni, and K. F. J. J. o. S. S. E. Chong,


"One-step electrochemical synthesis of MoS 2/graphene composite for
supercapacitor application," vol. 24, pp. 25-34, 2020.

[41] C. Koventhan et al., "Efficient hydrothermal synthesis of flake-like molybdenum


disulfide for selective electrochemical detection of metol in water real samples,"
vol. 15, no. 8, pp. 7390-7406, 2020.

[42] L. Sikong, M. Masae, K. Kooptarnond, W. Taweepreda, and F. J. A. s. s. Saito,


"Improvement of hydrophilic property of rubber dipping former surface with
Ni/B/TiO2 nano-composite film," vol. 258, no. 10, pp. 4436-4443, 2012.

[43] S. Aksay, T. Özer, and M. J. T. E. P. J.-A. P. Zor, "Vibrational and X-ray diffraction
spectra of SnS film deposited by chemical bath deposition method," vol. 47, no. 3,
p. 30502, 2009.

[44] G. H. Priya et al., "Physico-Chemical Attributes and Photodegradation Assessment


of Crystal Violet Dye by Utilizing TiO2/Sn2S3 (Sn = 0.25, 0.50, 0.75 M)
Nanocomposite Prepared Via Hydrothermal Strategy," Journal of Cluster Science,
vol. 34, no. 6, pp. 3013-3029, 2023/11/01 2023.

[45] J. Huang, H. Zhang, Z. Chen, W. Xiao, and Y. J. J. o. M. S. M. i. E. Zhang,


"Undulated TiO2− x microtubes modified with multilayer MoS2 nanoflakes for
high-performance photocatalytic applications," vol. 33, no. 22, pp. 18083-18095,
2022.

52
[46] C. Ma et al., "Elucidating the Synergic Effect in Nanoscale MoS2/TiO2
Heterointerface for Na‐Ion Storage," vol. 9, no. 35, p. 2204837, 2022.

[47] Y. Lei, P. Guo, M. Jia, W. Wang, J. Liu, and J. J. J. o. M. S. Zhai, "One-step


photodeposition synthesis of TiO 2 nanobelts/MoS 2 quantum dots/rGO ternary
composite with remarkably enhanced photocatalytic activity," vol. 55, pp. 14773-
14786, 2020.

[48] X. Zhu, X. Liang, X. Fan, and X. J. R. a. Su, "Fabrication of flower-like MoS 2/TiO
2 hybrid as an anode material for lithium ion batteries," vol. 7, no. 61, pp. 38119-
38124, 2017.

[49] M. Khan, S. Hussain, N. Alwadai, M. Fatima, U. Shahzad, and M. J. O. Iqbal,


"Heterostructure of TiO2 and SnS for enhancing the structural, optical and
photovoltaic properties of solar cells," vol. 300, p. 171625, 2024.

[50] L. Zhao et al., "Sandwich-like C@ SnS@ TiO2 anodes with high power and long
cycle for Li-ion storage," vol. 12, no. 5, pp. 5857-5865, 2020.

[51] B. Feng, C. Liu, W. Yan, J. Geng, and G. J. R. a. Wang, "MoS 2 nanotubes loaded
with TiO 2 nanoparticles for enhanced electrocatalytic hydrogen evolution," vol. 9,
no. 45, pp. 26487-26494, 2019.

[52] M. Zhang et al., "Enhanced NH3 Sensing Performance of Mo Cluster-MoS2


Nanocomposite Thin Films via the Sulfurization of Mo6 Cluster Iodides
Precursor," vol. 13, no. 3, p. 478, 2023.

[53] S. V. Kite et al., "Nanostructured TiO2 sensitized with MoS2 nanoflowers for
enhanced photodegradation efficiency toward methyl orange," vol. 6, no. 26, pp.
17071-17085, 2021.

[54] H. Fan, R. Wu, H. Liu, X. Yang, Y. Sun, and C. J. J. o. m. s. Chen, "Synthesis of


metal-phase-assisted 1T@ 2H-MoS 2 nanosheet-coated black TiO 2 spheres with
visible light photocatalytic activities," vol. 53, pp. 10302-10312, 2018.

53
[55] P. Zhao et al., "One-dimensional MoS2-decorated TiO2 nanotube gas sensors for
efficient alcohol sensing," vol. 674, pp. 252-258, 2016.

[56] Y. Luo et al., "3D carbon-coated stannous sulfide-molybdenum disulfide anodes


for advanced lithium-ion batteries," vol. 1, no. 7, pp. 2323-2331, 2020.

[57] J. Ru et al., "Covalent assembly of MoS2 nanosheets with SnS nanodots as linkages
for lithium/sodium‐ion batteries," vol. 132, no. 34, pp. 14729-14735, 2020.

[58] X. Xiao, Y. Wang, X. Xu, T. Yang, and D. J. M. C. Zhang, "Preparation of the


flower-like MoS2/SnS2 heterojunction as an efficient electrocatalyst for hydrogen
evolution reaction," vol. 487, p. 110890, 2020.

[59] X. Wang et al., "High supercapacitor and adsorption behaviors of flower-like MoS
2 nanostructures," vol. 2, no. 38, pp. 15958-15963, 2014.

[60] A. Numan, Y. Zhan, M. Khalid, and M. Hatamvand, "Introduction to


supercapattery," in Advances in Supercapacitor and Supercapattery: Elsevier,
2021, pp. 45-61.

[61] R. Barik, N. Devi, V. K. Perla, S. K. Ghosh, and K. J. A. S. S. Mallick, "Stannous


sulfide nanoparticles for supercapacitor application," vol. 472, pp. 112-117, 2019.

[62] B. Xie et al., "Hydrothermal synthesis of layered molybdenum sulfide/N-doped


graphene hybrid with enhanced supercapacitor performance," vol. 99, pp. 35-42,
2016.

[63] M. R. Pallavolu, Y. A. Kumar, G. Mani, R. A. Alshgari, M. Ouladsmane, and S.


W. J. J. o. E. C. Joo, "Facile fabrication of novel heterostructured tin disulfide
(SnS2)/tin sulfide (SnS)/N-CNO composite with improved energy storage capacity
for high-performance supercapacitors," vol. 899, p. 115695, 2021.

54

You might also like